Biogeochemical cycle

Source: Wikipedia, the free encyclopedia.

A biogeochemical cycle, or more generally a cycle of matter,

abiotic compartments of Earth. The biotic compartment is the biosphere and the abiotic compartments are the atmosphere, lithosphere and hydrosphere
.

For example, in the carbon cycle, atmospheric

and other nutrients.

There are biogeochemical cycles for many other elements, such as for oxygen, hydrogen, phosphorus, calcium, iron, sulfur, mercury and selenium. There are also cycles for molecules, such as water and silica. In addition there are macroscopic cycles such as the rock cycle, and human-induced cycles for synthetic compounds such as for polychlorinated biphenyls (PCBs). In some cycles there are geological reservoirs where substances can remain or be sequestered for long periods of time.

Biogeochemical cycles involve the interaction of biological, geological, and chemical processes. Biological processes include the influence of

metabolic processes essential for the cycling of nutrients and chemicals throughout global ecosystems. Without microorganisms many of these processes would not occur, with significant impact on the functioning of land and ocean ecosystems and the planet's biogeochemical cycles as a whole. Changes to cycles can impact human health. The cycles are interconnected and play important roles regulating climate, supporting the growth of plants, phytoplankton
and other organisms, and maintaining the health of ecosystems generally. Human activities such as burning fossil fuels and using large amounts of fertilizer can disrupt cycles, contributing to climate change, pollution, and other environmental problems.

Overview

Generalized biogeochemical cycle[2]
Simplified version of the nitrogen cycle

Energy flows directionally through ecosystems, entering as sunlight (or inorganic molecules for

continental plates, all play a role in this recycling of materials. Because geology and chemistry have major roles in the study of this process, the recycling of inorganic matter between living organisms and their environment is called a biogeochemical cycle.[3]

The six aforementioned elements are used by organisms in a variety of ways. Hydrogen and oxygen are found in water and

organic molecules, both of which are essential to life. Carbon is found in all organic molecules, whereas nitrogen is an important component of nucleic acids and proteins. Phosphorus is used to make nucleic acids and the phospholipids that comprise biological membranes. Sulfur is critical to the three-dimensional shape of proteins. The cycling of these elements is interconnected. For example, the movement of water is critical for leaching sulfur and phosphorus into rivers which can then flow into oceans. Minerals cycle through the biosphere between the biotic and abiotic components and from one organism to another.[4]

Ecological systems (ecosystems) have many biogeochemical cycles operating as a part of the system, for example, the water cycle, the carbon cycle, the nitrogen cycle, etc. All chemical elements occurring in organisms are part of biogeochemical cycles. In addition to being a part of living organisms, these chemical elements also cycle through abiotic factors of ecosystems such as water (hydrosphere), land (lithosphere), and/or the air (atmosphere).[5]

The living factors of the planet can be referred to collectively as the biosphere. All the nutrients — such as carbon, nitrogen, oxygen, phosphorus, and sulfur — used in ecosystems by living organisms are a part of a closed system; therefore, these chemicals are recycled instead of being lost and replenished constantly such as in an open system.[5]

The major parts of the biosphere are connected by the flow of chemical elements and compounds in biogeochemical cycles. In many of these cycles, the

biota plays an important role. Matter from the Earth's interior is released by volcanoes. The atmosphere exchanges some compounds and elements rapidly with the biota and oceans. Exchanges of materials between rocks, soils, and the oceans are generally slower by comparison.[2]

The flow of energy in an ecosystem is an open system; the Sun constantly gives the planet energy in the form of light while it is eventually used and lost in the form of heat throughout the trophic levels of a food web. Carbon is used to make carbohydrates, fats, and proteins, the major sources of food energy. These compounds are oxidized to release carbon dioxide, which can be captured by plants to make organic compounds. The chemical reaction is powered by the light energy of sunshine.

Sunlight is required to combine carbon with hydrogen and oxygen into an energy source, but ecosystems in the

oxidation and reduction of sulfur compounds (e.g., oxidizing elemental sulfur to sulfite and then to sulfate
).

  • Examples of major biogeochemical processes
    Examples of major biogeochemical processes
  • The oceanic whale pump showing how whales cycle nutrients through the ocean water column
    The oceanic
    whale pump showing how whales cycle nutrients through the ocean water column
  • The implications of shifts in the global carbon cycle due to human activity are concerning scientists.[6]
    The implications of shifts in the
    global carbon cycle due to human activity are concerning scientists.[6]

Although the Earth constantly receives energy from the Sun, its chemical composition is essentially fixed, as the additional matter is only occasionally added by meteorites. Because this chemical composition is not replenished like energy, all processes that depend on these chemicals must be recycled. These cycles include both the living biosphere and the nonliving lithosphere, atmosphere, and hydrosphere.

Biogeochemical cycles can be contrasted with geochemical cycles. The latter deals only with crustal and subcrustal reservoirs even though some process from both overlap.

Compartments

Atmosphere

Hydrosphere

Beach scene simultaneously showing the atmosphere (air), hydrosphere (ocean) and lithosphere (ground)
Some roles of marine organisms in biogeochemical cycling in the Southern Ocean[7]

The global ocean covers more than 70% of the Earth's surface and is remarkably heterogeneous. Marine productive areas, and

microbial assemblages at an unprecedented rate.[14][15][16][17][9]

Global change is, therefore, affecting key processes including

planktonic communities, principally through effects on calcifying taxa.[18] There is also evidence for shifts in the production of key intermediary volatile products, some of which have marked greenhouse effects (e.g., N2O and CH4, reviewed by Breitburg in 2018,[15] due to the increase in global temperature, ocean stratification and deoxygenation, driving as much as 25 to 50% of nitrogen loss from the ocean to the atmosphere in the so-called oxygen minimum zones[19] or anoxic marine zones,[20] driven by microbial processes. Other products, that are typically toxic for the marine nekton, including reduced sulfur species such as H2S, have a negative impact for marine resources like fisheries and coastal aquaculture. While global change has accelerated, there has been a parallel increase in awareness of the complexity of marine ecosystems, and especially the fundamental role of microbes as drivers of ecosystem functioning.[16][9]

Lithosphere

Biosphere

Microorganisms drive much of the biogeochemical cycling in the earth system.[21][22]

Reservoirs

The chemicals are sometimes held for long periods of time in one place. This place is called a reservoir, which, for example, includes such things as coal deposits that are storing carbon for a long period of time.[23] When chemicals are held for only short periods of time, they are being held in exchange pools. Examples of exchange pools include plants and animals.[23]

Plants and animals utilize carbon to produce carbohydrates, fats, and proteins, which can then be used to build their internal structures or to obtain energy. Plants and animals temporarily use carbon in their systems and then release it back into the air or surrounding medium. Generally, reservoirs are abiotic factors whereas exchange pools are biotic factors. Carbon is held for a relatively short time in plants and animals in comparison to coal deposits. The amount of time that a chemical is held in one place is called its

turnover time (also called the renewal time or exit age).[23]

Box models

Basic one-box model

Box models are widely used to model biogeochemical systems.[24][25] Box models are simplified versions of complex systems, reducing them to boxes (or storage reservoirs) for chemical materials, linked by material fluxes (flows). Simple box models have a small number of boxes with properties, such as volume, that do not change with time. The boxes are assumed to behave as if they were mixed homogeneously.[25] These models are often used to derive analytical formulas describing the dynamics and steady-state abundance of the chemical species involved.

The diagram at the right shows a basic one-box model. The reservoir contains the amount of material M under consideration, as defined by chemical, physical or biological properties. The source Q is the flux of material into the reservoir, and the sink S is the flux of material out of the reservoir. The budget is the check and balance of the sources and sinks affecting material turnover in a reservoir. The reservoir is in a steady state if Q = S, that is, if the sources balance the sinks and there is no change over time.[25]

The residence or turnover time is the average time material spends resident in the reservoir. If the reservoir is in a steady state, this is the same as the time it takes to fill or drain the reservoir. Thus, if τ is the turnover time, then τ = M/S.[25] The equation describing the rate of change of content in a reservoir is

When two or more reservoirs are connected, the material can be regarded as cycling between the reservoirs, and there can be predictable patterns to the cyclic flow.[25] More complex multibox models are usually solved using numerical techniques.

Simple three box model. Simplified budget of ocean carbon flows[26]
Measurement units

Global biogeochemical box models usually measure:

  • reservoir masses in
    petagrams
    (Pg)
  • flow fluxes in petagrams per year (Pg yr−1)

The diagram on the left shows a simplified budget of ocean carbon flows. It is composed of three simple interconnected box models, one for the

particles (marine snow) settle through the ocean interior. Only 2 Pg eventually arrives at the seafloor, while the other 8 Pg is respired in the dark ocean. In sediments, the time scale available for degradation increases by orders of magnitude with the result that 90% of the organic carbon delivered is degraded and only 0.2 Pg C yr−1 is eventually buried and transferred from the biosphere to the geosphere.[26]

More complex model with many interacting boxes. Export and burial rates of terrestrial organic carbon in the ocean[27]

The diagram on the right shows a more complex model with many interacting boxes. Reservoir masses here represents carbon stocks, measured in Pg C. Carbon exchange fluxes, measured in Pg C yr−1, occur between the atmosphere and its two major sinks, the land and the ocean. The black numbers and arrows indicate the reservoir mass and exchange fluxes estimated for the year 1750, just before the Industrial Revolution. The red arrows (and associated numbers) indicate the annual flux changes due to anthropogenic activities, averaged over the 2000–2009 time period. They represent how the carbon cycle has changed since 1750. Red numbers in the reservoirs represent the cumulative changes in anthropogenic carbon since the start of the Industrial Period, 1750–2011.[28][29][27]

Fast and slow cycles

The fast cycle operates through the biosphere, including exchanges between land, atmosphere, and oceans. The yellow numbers are natural fluxes of carbon in billions of tons (gigatons) per year. Red are human contributions and white are stored carbon.[30]
The slow cycle operates through rocks, including volcanic and tectonic activity

There are fast and slow biogeochemical cycles. Fast cycle operate in the biosphere and slow cycles operate in rocks. Fast or biological cycles can complete within years, moving substances from atmosphere to biosphere, then back to the atmosphere. Slow or geological cycles can take millions of years to complete, moving substances through the Earth's crust between rocks, soil, ocean and atmosphere.[31]

As an example, the fast carbon cycle is illustrated in the diagram below on the left. This cycle involves relatively short-term

seafloor sediments. The fast cycle includes annual cycles involving photosynthesis and decadal cycles involving vegetative growth and decomposition. The reactions of the fast carbon cycle to human activities will determine many of the more immediate impacts of climate change.[32][33][34][35]

The slow cycle is illustrated in the diagram above on the right. It involves medium to long-term

Mountain building processes result in the return of this geologic carbon to the Earth's surface. There the rocks are weathered and carbon is returned to the atmosphere by degassing and to the ocean by rivers. Other geologic carbon returns to the ocean through the hydrothermal emission of calcium ions. In a given year between 10 and 100 million tonnes of carbon moves around this slow cycle. This includes volcanoes returning geologic carbon directly to the atmosphere in the form of carbon dioxide. However, this is less than one percent of the carbon dioxide put into the atmosphere by burning fossil fuels.[31][32]

Deep cycles

The terrestrial subsurface is the largest reservoir of carbon on earth, containing 14–135

consortia[39][40][41] and small-scale metagenomic analyses of natural communities[42][43][44] suggest that organisms are linked via metabolic handoffs: the transfer of redox reaction products of one organism to another. However, no complex environments have been dissected completely enough to resolve the metabolic interaction networks that underpin them. This restricts the ability of biogeochemical models to capture key aspects of the carbon and other nutrient cycles.[45] New approaches such as genome-resolved metagenomics, an approach that can yield a comprehensive set of draft and even complete genomes for organisms without the requirement for laboratory isolation[42][46][47] have the potential to provide this critical level of understanding of biogeochemical processes.[48]

Some examples

Some of the more well-known biogeochemical cycles are shown below:

Many biogeochemical cycles are currently being studied for the first time. Climate change and human impacts are drastically changing the speed, intensity, and balance of these relatively unknown cycles, which include:

Biogeochemical cycles always involve active equilibrium states: a balance in the cycling of the element between compartments. However, overall balance may involve compartments distributed on a global scale.

As biogeochemical cycles describe the movements of substances on the entire globe, the study of these is inherently multidisciplinary. The carbon cycle may be related to research in

atmospheric sciences.[53] Biochemical dynamics would also be related to the fields of geology and pedology.[54]

See also

References

  1. ^ "CK12-Foundation". flexbooks.ck12.org. Retrieved 2022-03-21.
  2. ^ a b Moses, M. (2012) Biogeochemical cycles Archived 2021-11-22 at the Wayback Machine. Encyclopedia of Earth.
  3. ^ Biogeochemical Cycles Archived 2021-09-27 at the Wayback Machine, OpenStax, 9 May 2019. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine.
  4. ^ Fisher M. R. (Ed.) (2019) Environmental Biology, 3.2 Biogeochemical Cycles Archived 2021-09-27 at the Wayback Machine, OpenStax. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine.
  5. ^ a b "Biogeochemical Cycles". The Environmental Literacy Council. Archived from the original on 30 April 2015. Retrieved 20 November 2017.
  6. doi:10.1186/s13021-017-0077-x. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License. Archived 2017-10-16 at the Wayback Machine
    .
  7. hdl:11336/128446. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  8. .
  9. ^
    doi:10.3389/fmars.2019.00657. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  10. ^ Galton, D. (1884) 10th Meeting: report of the royal commission on metropolitan sewage Archived 2021-09-24 at the Wayback Machine. J. Soc. Arts, 33: 290.
  11. JSTOR 1294478
    .
  12. .
  13. .
  14. .
  15. ^ .
  16. ^ .
  17. ^ .
  18. .
  19. .
  20. .
  21. .
  22. PMID 33173062. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  23. ^ a b c Baedke, Steve J.; Fichter, Lynn S. "Biogeochemical Cycles: Carbon Cycle". Supplemental Lecture Notes for Geol 398. James Madison University. Archived from the original on 1 December 2017. Retrieved 20 November 2017.
  24. S2CID 4312683
    .
  25. ^ .
  26. ^
    doi:10.1007/978-3-030-10822-9. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  27. ^
    S2CID 30408500. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  28. .
  29. .
  30. ^ Riebeek, Holli (16 June 2011). "The Carbon Cycle". Earth Observatory. NASA. Archived from the original on 5 March 2016. Retrieved 5 April 2018.
  31. ^ .
  32. ^ from the original on 2021-09-27. Retrieved 2021-09-27.
  33. .
  34. doi:10.3390/sci1010017. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  35. from the original on 2021-11-22. Retrieved 2021-09-27.
  36. .
  37. .
  38. .
  39. .
  40. .
  41. .
  42. ^ from the original on 2021-09-27. Retrieved 2021-09-27.
  43. .
  44. .
  45. .
  46. .
  47. .
  48. PMID 27774985. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
    .
  49. ^ "Mercury Cycling in the Environment". Wisconsin Water Science Center. United States Geological Survey. 10 January 2013. Archived from the original on 11 April 2021. Retrieved 20 November 2017.
  50. .
  51. .
  52. .
  53. ^ McGuire, 1A. D.; Lukina, N. V. (2007). "Biogeochemical cycles" (PDF). In Groisman, P.; Bartalev, S. A.; NEESPI Science Plan Development Team (eds.). Northern Eurasia earth science partnership initiative (NEESPI), Science plan overview. Global Planetary Change. Vol. 56. pp. 215–234. Archived (PDF) from the original on 5 March 2016. Retrieved 20 November 2017.{{cite book}}: CS1 maint: numeric names: authors list (link)
  54. ^ "Distributed Active Archive Center for Biogeochemical Dynamics". daac.ornl.gov. Oak Ridge National Laboratory. Archived from the original on 11 February 2011. Retrieved 20 November 2017.

Further reading